Construction of Isochromenes via a Ruthenium ... - ACS Publications

On the basis of all of the information mentioned above, a possible mechanistic path accounting for the formation of the three products is pictured in ...
0 downloads 0 Views 255KB Size
6082

Organometallics 2007, 26, 6082-6090

Construction of Isochromenes via a Ruthenium-Catalyzed Reaction of Oxabenzonorbornenes with Propargylic Alcohols Karine Villeneuve and William Tam* Guelph-Waterloo Centre for Graduate Work in Chemistry and Biochemistry, Department of Chemistry, UniVersity of Guelph, Guelph, Ontario N1G 2W1, Canada ReceiVed May 7, 2007

The ruthenium-catalyzed cyclization of a propargylic alcohol with an oxabenzonorbornene in methanol leads to the creation of an isochromene framework. The proposed mechanism herein discussed for the formation of the product involves six major steps, the first four being oxidative cyclization, β-hydride elimination, hydroruthenation, followed by [2+2] cycloreversion. The ruthenium carbene formed at this stage undergoes a [1,3]-alkoxide shift that provides the observed product after reductive elimination. This process, believed to occur via a cationic ruthenium species, is in competition with two other pathways, ruthenium-catalyzed [2+2] cycloaddition and cyclopropanation. Although both [Cp*Ru(CH3CN)3]PF6 and Cp*Ru(COD)Cl are effective catalysts, the latter gives better yield and product ratio. The reaction was also found to proceed with high regioselectivity and product selectivity when unsymmetrical alkenes bearing a coordinating functional group at the bridge junction were used. Introduction Oxabicyclic alkenes are valuable synthetic intermediates as they can serve as useful building blocks in molecular architecture1 or as a general template to create highly substituted rings. For instance, transition metal-catalyzed asymmetric ring opening2 of these alkenes allows the formation of several stereocenters in a single step. Numerous transition metals have been explored, such as rhodium,3 palladium,4 and nickel.5 In our research program investigating ruthenium-catalyzed reactions, we have recently examined different aspects involving oxabenzonorbornenes 2.6 Ruthenium catalysts are known for their chemical transformation versatility,7,8 but they can be highly * Corresponding author. E-mail: [email protected]. (1) For selected examples of molecular architecture, see: (a) Girreser, U.; Giuffrida, D.; Kohnke, F. H.; Mathias, J. P.; Philp, D.; Stoddart, J. F. Pure Appl. Chem. 1993, 65, 119. (b) Warrener, R. N.; Margetic, D.; Sun, G.; Amarasekara, A. S.; Foley, P.; Butler, D. N.; Russell, R. A. Tetrahedron Lett. 1999, 40, 4111. (c) Warrener, R. N.; Butler, D. N.; Margetic, D.; Pfeffer, F. M.; Russell, R. A. Tetrahedron Lett. 2000, 41, 4671. (d) Warrener, R. N.; Margetic, D.; Foley, P.; Butler, D. N.; Winling, A.; Beales, K. A.; Russell, R. A. Tetrahedron 2001, 57, 571. (e) Dalphond, J.; Bazzi, H. S.; Kahrim, K.; Sleiman, H. F. Macromol. Chem. Phys. 2002, 203, 1988. (f) Warrener, R. N. Eur. J. Org. Chem. 2000, 3363. (2) Lautens, M.; Fagnou, K.; Heibert, S. Acc. Chem. Res. 2003, 36, 48. (3) Lautens, M.; Fagnou, K.; Yang, D. J. Am. Chem. Soc. 2003, 125, 14884 and references therein. (4) (a) Lautens, M.; Hiebert, S.; Renaud, J.-L. J. Am. Chem. Soc. 2001, 123, 6834. (b) Lautens, M.; Hiebert, S. J. Am. Chem. Soc. 2004, 126, 1437. (c) Cabrera, S.; Arraya´s, R. G.; Alonso, I.; Carretero, J. C. J. Am. Chem. Soc. Chem. 2005, 127, 17938. (d) Li, M.; Yan, X.-X.; Zhu, X.-Z.; Cao, B.-X.; Sun, J.; Hou, X.-L. Org. Lett. 2004, 6, 2833 and references therein. (5) Wu, M.-S.; Jeganmohan, M.; Cheng, C.-H. J. Org. Chem. 2005, 70, 9545. (6) (a) Villeneuve, K.; Jordan, R. W.; Tam, T. Synlett 2003, 2123. (b) Villeneuve, K.; Tam, W. Organometallics 2006, 25, 843. (c) Villeneuve, K.; Tam, W. J. Am. Chem. Soc. 2006, 128, 3514. (7) For a review on the use of Cp*Ru(COD)Cl, see: De´rien, S.; Dixneuf, P. H. J. Organomet. Chem. 2004, 689, 1382. (8) For a general review on ruthenium-catalyzed reactions, see: (a) Trost, B. M.; Toste, F. D.; Pinkerton, A. B. Chem. ReV. 2001, 101, 2067. (b) Naota, T.; Takaya, H.; Murahashi, S.-I. Chem. ReV. 1998, 98, 2599. (c) Topics in Organometallic Chemistry; Bruneau, C., Dixneuf, P. H., Eds.; Springer-Verlag GmbH: Berlin, New York, 2004; Vol. 11. (d) Ruthenium in Organic Synthesis; Murahashi, S.-I., Ed.; Wiley-VCH: Weinhein, 2004.

substrate-, ligand-, and solvent-dependent.9 We found that, by varying the reaction conditions, several products could be obtained (Scheme 1). For example, when alkene 2a is treated with Cp*Ru(COD)Cl (1a) (Cp* ) pentamethylcyclopentadienyl; COD ) 1,5-cyclooctadiene), isomerization to the corresponding naphthalene oxide 3 is observed.6c When an alkyne is added to the reaction mixture, Ru-catalyzed [2+2] cycloaddition usually occurs as the only pathway.6a,b However, when the alkyne is a secondary propargylic alcohol such as 4a, cyclopropane 6a is obtained in a significant amount or as the major product.6b Cyclopropane 6a is postulated to be generated through oxidative cyclization of the two unsaturated partners, followed by β-hydride elimination, hydroruthenation, and reductive elimination. It was shown that this pathway can, on the other hand, be mostly suppressed by utilizing Cp*Ru(COD)I (1b) as the precatalyst. More unexpectedly, when performing the same reaction in a protic solvent such as methanol, the formation of the isochromene product 7a is taking place.10 In this Article, we wish to provide more insight about the formation of isochromene 7a and the competition between these ruthenium-catalyzed processes.

Results and Discussion Ruthenium(II)-Catalyzed Formation of Isochromene. Isochromenes are a class of compounds that exhibit diverse biological activities,11 and the interest for their synthesis has recently exploded. The most common approach involves different variations of intramolecular cyclization of acetylenic aldehydes or ketones.12 Other methods also include Pd-catalyzed tandem reaction of pinacolone with aryl bromides,13 intramolecular carbolithiation of propargylic acetals,14 intramolecular (9) For selected examples, see: (a) Morisaki, Y.; Kondo, T.; Mitsudo, T. Organometallics 1999, 18, 4742. (b) Tanaka, D.; Sato, Y.; Mori, M. Organometallics 2006, 25, 799. (c) Trost, B. M.; Mu¨ller, T. J. J.; Martinez, J. J. Am. Chem. Soc. 1995, 117, 1888. (d) Yamamoto, Y.; Kitahara, H.; Ogawa, R.; Kawaguchi, H.; Tatsumi, K.; Itoh, K. J. Am. Chem. Soc. 2000, 122, 4310. (10) For some of the preliminary results of this work, see: Villeneuve, K.; Tam, W. Eur. J. Org. Chem. 2006, 5449.

10.1021/om7004518 CCC: $37.00 © 2007 American Chemical Society Publication on Web 11/07/2007

Construction of Isochromenes

Organometallics, Vol. 26, No. 25, 2007 6083

Scheme 1. Different Products Formed via Ruthenium-Catalyzed Reactions Involving Oxabicyclic Alkenes

Figure 1. GOESY experiment allowing one to determine the stereochemistry of the trisubstituted double bond in 7a.

istry of the trisubstituted double bond (Figure 1). Finally, the structure of 7a was unequivocally confirmed by X-ray analysis.18 Because the formation of isochromene 7a from alkene 2a and alkyne 4a could not be directly extrapolated, deuteration experiments were designed so that the change in atom connectivity from the starting materials to the product could be explained. Reaction of 2a-d4 with 4a produced the fully tetradeuterated isochromene 7a-d4 in a yield of 48% (eq 1).

iridium hydride-catalyzed alkyne hydroalkoxylation,15 Pdcatalyzed reaction of 2-iodophenyloxyallene with bicyclic alkenes,16 and 1,6-addition of nucleophiles to benzopyranylidenetungsten(0) complexes.17 When alkene 2a and alkyne 4a were reacted in MeOH in the presence of Cp*Ru(COD)Cl (1a), the new product, significantly less polar than 5a and 6a, was easily isolated as a sole isomer by column chromatography on silica. It displayed some spectral patterns similar to those observed for 6a. The molecular ion peak (M+, 286) indicated that 7a is a 1:1 adduct of 2a and 4a with no incorporation of methanol. However, in comparison to 6a, 1H and 13C (JMOD) NMR spectra of 7a clearly showed the loss of symmetry in the molecule, the disappearance of the bridged structure, and the presence of three vinylic protons. Further NMR experiments (HMBC, HSQC), IR spectroscopy, and mass spectrometry allowed one to elucidate the structure of 7a, and GOESY NMR experiments assigned the stereochem(11) (a) Lacassagne, A.; Buu-Hoi, N. P.; Zajdela, F.; Stora, F.; Mangane, M.; Jacquignon, P. C. R. Seances Acad. Sci., Ser. D 1971, 272, 3102. (b) Thines, E.; Anke, H.; Sterner, O. J. Nat. Prod. 1998, 61, 306. (c) Wang, W.; Li, T.; Milbum, R.; Yates, J.; Hinnant, E.; Luzzio, M. J.; Noble, S. A.; Attardo, G. Bioorg. Med. Chem. Lett. 1998, 8, 1579. (d) Poch, G. K.; Gloer, J. B. Tetrahedron Lett. 1989, 30, 3483. (e) Hari, L.; De Buyck, L. F.; Pooter, H. L. Phytochemistry 1991, 30, 1726. (12) (a) Asao, N.; Nogami, T.; Takahashi, K.; Yamamoto, Y. J. Am. Chem. Soc. 2002, 124, 764. (b) Mondal, S.; Nogami, T.; Asao, N.; Yamamoto, Y. J. Org. Chem. 2003, 68, 9496. (c) Barluenga, J.; Va´zquezVilla, H.; Ballesteros, A.; Gonza´lez, J. J. Am. Chem. Soc. 2003, 124, 9028. (d) Patil, N. T.; Yamamoto, Y. J. Org. Chem. 2004, 69, 5139. (e) Yue, D.; Della Ca´, N.; Larock, R. C. Org. Lett. 2004, 6, 1581. (f) Patil, N. T.; Pahadi, N. K.; Yamamoto, Y. J. Org. Chem. 2005, 70, 10096. (g) Yue, D.; Della Ca´, N.; Larock, R. C. J. Org. Chem. 2006, 71, 3381. (h) Yao, X.; Li, C.-J. Org. Lett. 2006, 8, 1953. (13) Mutter, R.; Campbell, I. B.; Martin de la Nava, E. M.; Merritt, A.; T.; Wills, M. J. Org. Chem. 2001, 66, 3284. (14) Le Strat, F.; Harrowven, D. C.; Maddaluno, J. J. Org. Chem. 2005, 70, 489. (15) Li, X.; Chianese, A. R.; Vogel, T.; Crabtree, R. H. Org. Lett. 2005, 7, 5437. (16) Parthasarathy, K.; Jeganmohan, M.; Cheng, C.-H. Org. Lett. 2006, 8, 621. (17) Kusama, H.; Sawada, T.; Okita, A.; Shiozawa, F.; Iwasawa, N. Org. Lett. 2006, 8, 1077.

When the propargylic position was deuterated (4a-d), the deuterium was found on the carbon adjacent to the ketone, as in cyclopropane 6a-d (Scheme 2). In this case, a significant empirical isotopic effect was observed as longer reaction time was needed and the 5a:6a:7a ratio changed from 20:6:74 with 4a to 41:10:49 with 4a-d. This clearly indicates that the ease of breaking the propargylic C-H bond is a determinant factor in the formation of 7a, which is very comparable to what we observed in our previous study of the formation of 6a.19 We thus hypothesized that the formation of 7a may be related to the formation of 6a (vide infra). On the other hand, when the reaction was performed in MeOD, no substantial difference in the products distribution was found. However, unlike the 100% deuterium incorporation obtained for the reaction with 4a-d, the reaction in MeOD produced 7a-d′ with 86% of Dincorporation. This is presumably due to an incomplete exchange of a labile proton in 4a or an intermediate with the solvent. It is also noteworthy that these two reactions produce the opposite major diastereomer (Scheme 3). Factors Influencing the Formation of Isochromene versus [2+2] Cycloadduct and Cyclopropane. To understand the formation of 7a, several reaction parameters were studied: the catalyst, the presence of additives, the solvent, and the substituents on the two reactive moieties. Catalyst Effect. While 1a efficiently catalyzes the formation of 7a, other Ru-complexes such as CpRu(COD)Cl, [CpRu(CH3CN)3]PF6, and CpRu(PPh3)2Cl were found to be inactive. This agrees with our previous observation for the formation of 5a and 6a, where the more electron-rich Cp*Ru(COD)Cl was required for the reaction to occur.6b However, when [Cp*Ru(CH3CN)3]PF6 (1c) was utilized in MeOH, a complex mixture with trace amounts of 5a, 6a, and 7a was obtained (entry 2, Table 1). (18) Lough, A. J.; Villeneuve, K.; Tam, W. Acta Crystallogr. 2006, E62, o2582. (19) In THF, the 5a:6a ratio for the reaction with 4a was 31:69, whereas 54:46 was obtained with 4a-d; see ref 6b.

6084 Organometallics, Vol. 26, No. 25, 2007

VilleneuVe and Tam

Scheme 2. Deuterium Labeling Experiments

Scheme 3. Partial 1H NMR Spectra of 7a, 7a-d, and 7a-d′

Intrigued by the fact that isochromene 7a is formed predominantly in MeOH (entry 1), whereas cyclopropane 6a is the major product and 7a is not observed in THF (entry 3), investigation of the nature of the active catalytic species was undertaken. Complex 1a has been postulated by Mitsudo and co-workers20 to form a neutral [Cp*RuCl] species for ruthenium-catalyzed [2+2] cycloaddition; on the other hand, it is believed that a cationic ruthenium species is formed in MeOH.21 To verify this, Cp*Ru(COD)Cl was treated with AgOTf in THF to form [Cp*Ru]+ prior to adding alkyne 4a and alkene 2a, and, indeed, (20) Mitsudo, T.; Naruse, H.; Kondo, T.; Ozaki, Y.; Watanabe, Y. Angew. Chem., Int. Ed. Engl. 1994, 31, 580. (21) Davies, S. G.; McNally, J. P.; Smallridge, A. J. AdV. Organomet. Chem. 1990, 30, 1.

isochromene 7a was obtained as the major product, although the yield was decreased (entry 4). A similar result was also found when [Cp*Ru(CH3CN)3]PF6 was utilized (entry 5). Thus, the formation of 7 appears to proceed via an active cationic ruthenium species, and a protic solvent is not necessarily required. Additive Effect. Attention was also drawn to the fact that cationic ruthenium complex [Cp*Ru(CH3CN)3]PF6 (1c) was giving a much more complex products mixture in MeOH (entry 2, Table 1) than in THF (entry 5, Table 1), whereas an overall yield for the three products of 69% was obtained with Cp*Ru(COD)Cl in MeOH (entry 1, Table 1). A closer look at the presence of the halide on the Ru-complex was then taken. Varying the concentration of chloride ions through addition of tetrabutylammonium chloride revealed the importance of the

Construction of Isochromenes

Organometallics, Vol. 26, No. 25, 2007 6085 Table 1. Effect of Catalytic Species on the Reaction Outcome

ratiob

yieldc (%)

entry

reaction conditionsa

5a:6a:7a

5a

6a

7a

1 2 3 4 5

1a/MeOH 1c/MeOH 1a/THF 1a/AgOTfd/THF 1c/THF

20:6:74 N/A 31:69:0 18:11:71 20:25:55

14 trace 25 8 11

4 trace 56 5 14

51 (50) trace 0 31 31

a Ru-complex (5 mol %): 1a ) Cp*Ru(COD)Cl, 1c ) [Cp*Ru(CH CN) ]PF . b Determined by analysis of the crude 1H NMR spectrum. c Yields were 3 3 6 based on the 1H NMR spectrum of the crude reaction mixture with mesitylene as internal standard. Yields in brackets are isolated yields. d 1 equiv with respect to the ruthenium precatalyst.

Table 2. Effect of Halide Additives on the Reaction Outcome

ratioc entry 1 2 3 4 5 6 7 8 9 10

precatalysta 1a 1a 1c 1c 1a 1c 1a 1b 1a 1b

ratioc

yieldd (%)

solvent

additiveb

5a:(6a+7a)

5a:6a:7a

5a

6a

7a

MeOH MeOH MeOH MeOH THF THF MeOH MeOH THF THF

none Bu4NCl(9)e none Bu4NCl(12) Bu4NCl(10) Bu4NCl(10) Bu4NI(10) none none none

20:80 20:80 N/A 16:84 N/A N/A 44:56 40:60 31:69 87:13

20:6:74 20:7:73 N/A 16:15:69 N/A N/A 44:5:51 40:2:58 31:69:0 87:13:0

14 17 trace 14 trace trace 36 35 25 55

4 6 trace 13 13 14 4 2 56 8

51 62 trace 59 0 0 41 51 0 0

a Ru-complex (5 mol %): 1a ) Cp*Ru(COD)Cl, 1b ) Cp*Ru(COD)I, 1c ) [Cp*Ru(CH CN) ]PF . b Numbers in bracket are the number of equivalents 3 3 6 with respect to the ruthenium precatalyst. c Determined by analysis of the crude 1H NMR spectrum. d Yields were based on the 1H NMR spectrum of the crude reaction mixture with mesitylene as internal standard. e Similar result was obtained when utilizing Bu4PCl instead of Bu4NCl.

halide (TBAC, entries 1-10, Table 2). Although no significant difference was observed with Cp*Ru(COD)Cl (entry 2), adding TBAC significantly improved the overall yield of the reaction (entry 4) when cationic ruthenium complex 1c was used as precatalyst.22 In contrast, the presence of a large excess of TBAC in THF completely deactivates both 1a and 1c precatalysts (entries 5 and 6). Furthermore, similarly to what was previously shown for the formation of the cyclopropane 6a,6b varying the halide on the catalyst drastically influences the product distribution, as the formation of the [2+2] cycloadduct 5a over cyclopropane 6a was following the trend I > Br > Cl. As expected, performing the reaction in the presence of 1a with tetrabutylammonium iodide or with Cp*Ru(COD)I (1b) increased the formation of 5a versus 6a and 7a (compare entries 7 and 8 with 1, Table 2). However, this change in the 5a:(6a+7a) ratio is not as important as what we observed in THF (compare entries 8 with 1 and 10 with 9). Along with halide-based salts, other additives were tested. However, none of these proved effective. The presence of the (22) Such chloride ion effects have been previously reported for CpRu(COD)Cl, see: Trost, B. M.; Indolese, A. F.; Mu¨ller, T. J. J.; Treptow, B. J. Am. Chem. Soc. 1995, 117, 615.

strongly coordinating ligand triphenylphosphine completely deactivates the catalyst,23 and no reaction was found to occur. The use of acid such as p-toluenesulfonic acid24 or strong chloride scavenger like silver triflate25 produces complex products mixture with trace amounts of products. The weaker chloride scavenger NH4PF69a,22,26 was compatible with the reaction conditions, although no yield or ratio improvement was obtained. Solvent Effect. It is evident that solvent plays a crucial role in the formation of 7a. The preceding study6b of the formation of 6a versus 5a has shown only a minor difference in the product distribution when the polarity of the aprotic solvent was varied. To further confirm that the polarity is not playing a considerable (23) Strongly complexing molecules are believed to prevent the crucial oxidative cyclization step, see: (a) Yamamoto, Y.; Nakagai, Y.; Itoh, K. Chem.-Eur. J. 2004, 10, 231. (b) Reference 21. (24) (a) Trost, B. M.; Dyker, G.; Kulawiec, R. J. J. Am. Chem. Soc. 1990, 112, 7809. (b) Trost, B. M.; Brown, R. E.; Toste, F. D. J. Am. Chem. Soc. 2000, 122, 5877. (25) Trost, B. M.; Probst, G. D.; Schoop, A. J. Am. Chem. Soc. 1998, 120, 9928. (26) Yamamoto, Y.; Kitahara, H.; Hattori, R.; Itoh, K. Organometallics 1998, 17, 1910.

6086 Organometallics, Vol. 26, No. 25, 2007

VilleneuVe and Tam Table 3. Solvent Effect

ratiob entry

catalytic systema

solvent

1 2 3 4 5 6 7 8

1a 1a/AgOTf 1c 1a 1a 1a 1a 1a

CH3CN CH3CN CH3CN TFEOH MeOH EtOH n-PrOH i-PrOH

ratiob

ratiob

yieldc (%)

5a:(6a+7a)

6a:7a

5a:6a:7a

5a

6a

7a

29:71 29:71 28:72 68:32 20:80 21:79 22:78 28:72

100:0 96:4 92:8 6:94 8:92 9:90 24:76 50:50

29:71:0 29:68:3 28:66:6 68:2:30 20:6:74 21:8:71 22:19:59 28:36:36

19 25 23 34 14 17 15 20

46 59 53 1 4 7 13 26

0 3 5 15 51 58 41 26

a Ru-complex (5 mol %): 1a ) Cp*Ru(COD)Cl, 1c ) [Cp*Ru(CH CN) ]PF . b Determined by analysis of the crude 1H NMR spectrum. c Yields were 3 3 6 based on the 1H NMR spectrum of the crude reaction mixture with mesitylene as internal standard.

role in the formation of 7a, the reaction of 2a and 4a in presence of Cp*Ru(COD)Cl was performed in acetonitrile, whose dielectric constant is comparable to methanol (entry 1, Table 3). Not surprisingly, no isochromene was produced, and a 5a: 6a ratio of 29:71 was observed. However, when the cationic ruthenium complex 1c was utilized instead of 1a (entries 2 and 3), very little 7a was generated and cyclopropane 6a was the major product formed. This is especially different from what is observed in THF (entries 4 and 5, Table 1), which suggests that the way the solvent acts as a ligand on the ruthenium may prevent or favor the formation of one product versus another.27 Consequently, the active intermediate in the formation of 6a is also probably a cationic ruthenium species. Other protic solvents were then investigated (entries 4-8, Table 3). With the exception of trifluoroethanol, no difference in the 5a:(6a+7a) ratio was observed. In contrast, a much broader variation of the 6a:7a ratio was found, especially in the case of n-propanol and isopropanol. Substrate Scope. Table 4 presents several alkyne and alkene partners that were subjected to the reaction conditions with the precatalyst 1a. Unlike secondary alcohol 4a, alkynes 4b (primary alcohol) and 4c (tertiary alcohol) (entries 2 and 3) only generated the [2+2] cycloadduct. This further suggests that isochromenes are formed through a reaction path similar to the one for cyclopropanes 6, because the same trend (i.e., only 5 was produced) was observed when the reaction was performed in THF.6b Other secondary propargylic alcohols such as those described in entries 4-9 were also tested. Steric hindrance at the propargylic position appears to play a dual role in the product distribution. A small increase of the steric hindrance does not affect the outcome of the reaction (compare entries 1 and 4), but when the alcohol side of the alkyne becomes too bulky, the [2+2] cycloadduct predominates. This is demonstrated with alkyne 4e (entry 5), bearing a tert-butyl group at the propargylic position, from which 5e was isolated in a yield of 54%. On the other hand, utilizing alkyne such as 4f (entry 6) seems beneficial for the formation of the isochromene product, and 7g was isolated in 70% yield. In contrast, the presence of a phenyl group (entries 7 and 8) somewhat appreciably decreases the yield. The other side of the alkyne was also modified, and it was found (27) Similar argument has been presented for the use of acetonitrile in other Ru-catalyzed processes, see: Kondo, T.; Morisaki, Y.; Uenoyama, S.; Wada, K.; Mitsudo, T. J. Am. Chem. Soc. 1999, 121, 8657.

that substituting the ethyl for the bulkier tert-butyl ester greatly enhances the formation of the [2+2] cycloadduct 5i over the isochromene 7i (compare entries 1 and 9). As for the olefin moiety, other symmetrical oxabenzonorbornenes (2b and 2c) were also exposed to the reaction conditions with 4a and were found to display reactivity comparable to 2a (entries 10 and 11, Table 4). In opposition, oxanorbornene 2d and oxanorbornadiene 2e (entries 12 and 13), lacking the aromatic ring, did not provide the rearranged product but rather the cyclopropane. This establishes the importance of the ring strain in the formation of 7. The use of 2f, bearing a carbon-based bridgehead, prevents the pathway leading to the formation of 6 and 7 (entry 14). Unsymmetrical alkenes were also utilized, and it was found that the presence and the nature of a group at the bridge junction critically modify the outcome of the reaction. The methyl group that bears alkene 2g increases the steric congestion at the reaction center, and cycloadduct 5o was formed as the major product (entry 15). Alternatively, the presence of a polar group at the bridge junction such as methyl ester (2h, entry 16) or methyl ketone (2i, entry 17) appears to promote the formation of the isochromene, and 7p and 7q were, respectively, generated as sole products. Hence, the presence on the olefin of a group capable of coordinating to the Rumetal can dictate the regiochemistry, but most importantly works synergistically with the propargylic alcohol to yield exclusively the isochromene product. Finally, alkene 8 (eq 2), bearing two potentially reactive double bonds, was synthesized. Although non-strained alkenes are known to undergo Ru-catalyzed Alderene reaction with propargylic alcohols,9c,28 we were delighted to only isolate the highly functionalized isochromene 9 in 74% yield.

Reaction Mechanism. From the effects of reaction parameters, the focus was then turned on how the isochromenes 7 are (28) Trost, B. M.; Mu¨ller, T. J. J. J. Am. Chem. Soc. 1994, 116, 4985.

Construction of Isochromenes

Organometallics, Vol. 26, No. 25, 2007 6087 Table 4. Scope of the Reaction

entry

a

alkyne

alkene

1 2 3 4 5

4a 4b 4c 4d 4e

2a 2a 2a 2a 2a

6 7 8 9

4f 4g 4h 4i

2a 2a 2a 2a

10 11 12 13 14 15 16 17

4a 4a 4a 4a 4a 4a 4a 4a

2b 2c 2d 2e 2f 2g 2h 2i

isolated product (yieldb) 7a (50%) 5b (38%) 5c (57%) 7d (50%) 5e (54%) 7e (26%) 7f (70%) 7g (34%) 7h (33%) 5i (32%) 7i (37%) 7j (60%) 7k (52%) 6l (30%) 6m (77%) 5n (50%) 5o (53%) 7p (77%) 7q (63%)

Determined by analysis of the characteristic peaks in the crude 1H NMR spectrum. b Isolated yields after column chromatography.

formed. Although oxidative cyclization of the alkene and the alkyne moieties with the ruthenium catalyst would be the first obvious step in the formation of the isochromene, another pathway could not be ruled out yet. As mentioned above (Scheme 1), alkene 2a exhibits a competitive reactivity in the presence of the precatalyst 1a, where either opening of the oxygen-bridge or oxidative cyclization with an alkyne can occur. In addition, when 2a is treated with 1a in MeOH, racemic 1,2dihydro-2-methoxy-1-naphthalenol 10 is obtained in a yield of 66% (eq 3). However, the ring-opening pathway could be excluded by rationalizing the regiochemistry obtained in isochromene 7p (entry 16, Table 4). If the first step involves a pathway similar to that for the formation of 11 through the intermediate 13 (Scheme 4), isochromene 14 would be the expected product. On the other hand, Burton and Tam have recently shown that propargylic alcohol 4c reacts in a highly regioselective fashion with 2h to give almost exclusively cycloadduct 12.29 Consequently, the fact that 7p was isolated indicates the formation of intermediate 15 via oxidative cyclization.

although the formation of 10 is possible, it should be noted that this compound is produced only when an excess of 2a is used (eq 4). To better understand these competitive processes, the reaction was monitored directly by 1H NMR spectroscopy (Figure 2). We found that the formation of the nucleophilic ringopening product 10 did not occur until the concentration of the alkyne is low (T ) 20 min, Figure 2). This shows that if both the alkyne and the alkene are available to complex to the catalyst, oxidative cyclization occurs more rapidly, and, as a result, only excess oxabicyclic alkene is converted into 10.

On the basis of all of the information mentioned above, a possible mechanistic path accounting for the formation of the Thus, several reaction pathways are available under the actual reaction conditions (Cp*Ru(COD)Cl, MeOH, 60 °C), and

(29) Burton, R. R.; Tam, W. Tetrahedron Lett. 2006, 47, 7185.

6088 Organometallics, Vol. 26, No. 25, 2007

VilleneuVe and Tam

Scheme 4. Rationalizing the Oxidative Cyclization Step

three products is pictured in Scheme 5. After oxidative cyclization of 2a and 4a with 1a, four possible ruthenacycle intermediates can be formed (Scheme 6). Upon reductive elimination of any of them, cycloadduct 5a is generated. However, only 16 and 16′ can undergo further β-hydride elimination as the β-hydrogen is not accessible in 21 and 21′. This could explain why some [2+2] cycloadduct is present in most cases, and would be in coherence with previous work on Ru-catalyzed Alder-ene reaction of propargylic alcohols by Trost and coworkers.9c,28 They have shown that increasing steric hindrance at the propargylic position improves the ratio favoring 22 (Scheme 7). Also, although their precatalyst of choice was CpRu(COD)Cl, the 5a:(6a+7a) ratio of 20:80 (entry 5, Table 3) is very similar to the 22:23 ratio they observed with alkyne 4a (4.5:1). This would also be consistent with the 5a:(6a+7a) ratio of 68:32 obtained in trifluoroethanol (entry 4, Table 3), as they also notice a lower 22:23 ratio of 2.3:1 in this solvent. It was previously emphasized (Table 2) that the presence of a halide ion plays a dual role in the reaction process. Although the ionization of the halide is not necessary for the generation of [2+2] cycloadduct, it seems required for the formation of cyclopropane and isochromene products. On the other hand, the presence of the halide counteranion appears to be crucial for the products formation, especially in MeOH where only trace amounts of the products were found when [Cp*Ru(CH3CN)3]PF6 was used. The halide moiety is suspected to stabilize the cationic ruthenium intermediates involved in the catalytic cycle. Thus, we postulate that the observed difference in reactivity is due to this ion-pairing.30 Once ruthenacycles 16 and 16′ are produced, either reductive or β-hydride elimination can occur, and increasing the steric hindrance should favor the reductive elimination process (Scheme 5).31 Therefore, decreasing steric bulk of the halide (I > Br > Cl) is in good agreement with the above observation, where more β-hydride elimination was observed with X ) Cl than with X ) I. This would also explain (30) Ion-pairing is known to considerably affect chemical reactions, see: Macchioni, A. Chem. ReV. 2005, 105, 2039. (31) In our previous report on ruthenium-catalyzed cyclopropanation of oxabicyclic alkenes with propargylic alcohols, a 1,2-hydride migration was also postulated as another possible path. This is, however, probably not at play, because the variation of the products distribution with the solvent polarity is of small magnitude. This change is probably due to a better ionization of the chloride in more polar solvents.

Figure 2. Partial 400 MHz 1H NMR spectrum at 60 °C in MeOD of the crude reaction between 2a and 4a in the presence of 1a over time (2a:4a ) 1.1:1).

the fact that the halide effect is greater in aprotic than protic solvents, because ion-pairing is expected to be tighter in the first case. Decomplexation of the alcohol is also necessary to generate the proper orbital alignment for β-hydride elimination, and the presence of an oxygen at the bridgehead of the bicyclic framework may assist this departure. The allenol 17 can then tautomerize to the corresponding ketone prior to undergoing hydroruthenation and thus generate the ruthenacyclobutane 18. If this intermediate reductively eliminates, cyclopropane 6a is formed (path C). Alternatively, [2+2] cycloreversion of 18 would form the ruthenium carbene 19 (path D), which could rearrange to 20 through a 1,3-migration of the alkoxide group and finally reductively eliminate to produce the isochromene 7a.32 In addition to the similar isotopic effects observed in the formation of both products, the fact that the cis relationship of the ketone and the isochromene groups in 7a matches the stereochemistry in cyclopropane 6a (with the ketone group syn to the bicyclic structure) suggests that both products originate from the same intermediate 18 (Figure 3). Moreover, the isomer of isochromene 7a where these two groups are trans, like in the case of 6a,6b was not observed.

Figure 3. 5a and 7a possibly arise from the same intermediate 18. (32) Ruthenium carbene rearrangements have been proposed in other ruthenium-catalyzed reactions, see: (a) Reference 24a. (b) Trost, B. M.; Kulawiec, R. J. J. Am. Chem. Soc. 1992, 114, 5579.

Construction of Isochromenes

Organometallics, Vol. 26, No. 25, 2007 6089 Scheme 5. Plausible Mechanistic Pathway for the Formation of 7a

Scheme 6. Four Possible Ruthenacyclopentene Intermediates

Scheme 8. Ruthenium-Catalyzed Reaction of Enyne Described by Mori and Co-workers

Scheme 7. Alder-ene Reaction Reported by Trost and Co-workers

conditions produced only cyclopropane products 26 and 27 (Scheme 9). These results strongly support our mechanistic hypothesis and are also in agreement with the trend of reactivity shown in Table 4 for those two alkenes. Thus, the strain in the alkene moiety plays an important role in the formation of the isochromene over the cyclopropane product. Presumably, an (34) It was hoped that product A could be isolated. A possible reason for this unreactivity towards intramolecular cyclopropanation would be the formation of a chelated subtrate-carbene complex such as B that would prevent this pathway.

Utilizing the same precatalyst Cp*Ru(COD)Cl (1a), Mori and co-workers recently trapped similar ruthenium carbene (Scheme 8).9a,33 So far, all of our attempts to trap the intermediate 14 (Scheme 5 and eq 2) have failed.34 On the other hand, Dixneuf and co-workers have reported the generation of ruthenium carbene by reacting 1a with ethyl diazoacetate.35 On the basis of this work, if one could react 2a with a similar ruthenium carbene obtained from 1c and ethyl diazoacetate, the related ruthenacycle 24 would be formed, and the proposed formation of isochromene from ruthenacyclobutane intermediate 18 could be tested (Scheme 9). Indeed, when performing this reaction, isochromene 25 was obtained in 32% yield. Such ruthenium carbene reactivity is very different from that usually observed.36,37 In addition, subjecting 2d to the same reaction (33) Mori, M.; Saito, N.; Tanaka, D.; Takimoto, M.; Sato, Y. J. Am. Chem. Soc. 2003, 125, 5606.

Similar stable complexes have been previously described, see: (a) Grubbs, R. H.; Miller, S. J.; Fu, G. F. Acc. Chem. Res. 1995, 28, 446. (b) Feldman, J.; Murdzek, J. S.; Davis, W. M.; Schrock, R. R. Organometallics 1989, 8, 2260. (35) (a) Monnier, F.; Castillo, D.; De´rien, S.; Toupet, L.; Dixneuf, P. H. Angew. Chem., Int. Ed. 2003, 42, 5474. (b) Eckert, M.; Monnier, F.; Shchetnikov, G. T.; Titanyuk, I. D.; Osipov, S. N.; Toupet, L.; De´rien, S.; Dixneuf, P. H. Org. Lett. 2005, 7, 3741.

6090 Organometallics, Vol. 26, No. 25, 2007 Scheme 9. Reaction of a Ruthenium Carbene Generated from 1c and Ethyl Diazoacetate with Alkenes 2a and 2d

VilleneuVe and Tam

catalyst, followed by a β-hydride elimination, tautomerization, and hydroruthenation. The ruthenacyclobutane thus obtained further undergoes [2+2] cycloreversion to form a ruthenium carbene intermediate that rearranges via a [1,3]-alkoxide shift group and finally reductively eliminates to produce the desired compound. Cationic ruthenium intermediate is proposed to be the active catalyst species. This process is in competition at different stages of the catalytic cycle with two other reactions, Ru-catalyzed [2+2] cycloaddition and cyclopropanation of the alkyne with the alkene. Therefore, several factors need to be taken into consideration to influence the outcome of the reaction such as the nature (e.g., polarity, acidity, complexing abilities) of the solvent, the steric bulkiness of the propargylic alcohol, as well as the strain and the nature of the bridgehead group of the alkene. Depending on the reaction conditions, each compound can be obtained as the major product: Cp*Ru(COD)I in THF would generate the [2+2] cycloadduct 5, Cp*Ru(COD)Cl in CH3CN or THF would mainly give the cyclopropane 6 product, whereas Cp*Ru(COD)Cl in MeOH would produce the isochromene 7.

Experimental Section

increase of the strain in the ruthenacyclobutane 18 could favor the retro [2+2] cycloaddition over the reductive elimination step. As for the influence of the solvent, its mode of action in the [2+2] cycloreversion/reductive elimination dual of the ruthenacyclobutane intermediate still needs to be clarified.

Conclusion To summarize, we have found a Ru-catalyzed cyclization of oxabenzonorbornenes with propargylic alcohols. Although yields are moderate in some cases, this method allows the use of readily available starting materials for the construction of the synthetically useful isochromene skeleton. The catalytic cycle to form this product is believed to involve an oxidative cyclization of the two unsaturated partners with the ruthenium (36) In general, ruthenium carbenes undergo metathesis or cyclopropanation reaction. For reviews on ruthenium-catalyzed metathesis, see: (a) Grubbs, R. H.; Chang, S. Tetrahedron 1998, 54, 4413. (b) Armstrong, S. K. J. Chem. Soc., Perkin Trans. 1 1998, 371. (c) Fu¨rstner, A. Angew. Chem., Int. Ed. 2000, 39, 3013. For reviews covering cyclopropanation involving ruthenium carbenes, see ref 7. For other non-metathesis reactions, see: Alcaide, B.; Almendros, P. Chem.-Eur. J. 2003, 9, 1259. (37) It is noteworthy that alkene 2a is known to undergo ring-opening polymerization metathesis with catalysts such as [RuCl2(p-cymene)]2 (Delaude, L.; Demonceau, A.; Noels, A. F. Macromolecules 1999, 32, 2091) and RuCl2(CHPh)(PCy3)2 (Amir-Ebrahimi, V.; Corry, D. A.; Hamilton, J. G.; Thompson, J. M.; Rooney, J. J. Macromolecules 2000, 33, 717. AmirEbrahimi, V.; Rooney, J. J. J. Mol. Catal. A 2004, 212, 107).

A representative procedure, including characterization of the key product 7a, is described here. Full details are found in the Supporting Information. General Procedure for the Preparation of Isochromene Products. Isochromene 7a. A mixture of alkene 2a (205 mg, 1.42 mmol), acetylene 4a (180 mg, 1.27 mmol), and MeOH (1.6 mL) in an oven-dried vial was added via a cannula to an oven-dried screw-cap vial containing Cp*Ru(COD)Cl (weighed out from a dry box, 17 mg, 0.045 mmol) under nitrogen. The reaction mixture was stirred at 60 °C for 1 h. The solvent was evaporated, and the crude product was purified by column chromatography (gradient elution, EtOAc:hexanes ) 1:19 to 1:4) to provide 7a (181 mg, 0.633 mmol, 50%). Rf 0.33 (EtOAc/hexanes ) 1:4); IR (neat) 3070, 2985, 2955, 2936, 1720, 1713, 1622 cm-1; 1H NMR (CDCl3, 400 MHz) δ 7.22-7.26 (m, 2H), 7.17 (app t, 1H, J ) 7.5 Hz), 6.99 (app t, 2H, J ) 7.5 Hz), 6.53 (d, 1H, J ) 5.7 Hz), 5.84 (d, 1H, J ) 5.7 Hz), 5.73 (d, 1H, J ) 8.8 Hz), 4.18-4.24 (m, 2H), 3.63 (d, 1H, J ) 17.1 Hz), 3.53 (d, 1H, J ) 17.1 Hz), 2.23 (s, 3H), 1.29 (t, 3H, J ) 7.1 Hz); 13C NMR (APT, CDCl3, 100 MHz) δ 204.5, 166.5, 144.5, 139.5, 129.6, 129.4, 128.7, 128.0, 127.2, 124.4, 123.5, 105.5, 73.7, 61.4, 41.8, 29.9, 14.1. HRMS (EI) calcd for C17H18O4 (M+), 286.1205; found, 286.1207.

Acknowledgment. This work was supported by NSERC (Canada). K.V. would like to thank NSERC for providing a CGS-D2 scholarship. The authors would also like to acknowledge Gavin Chit Tsui and Ryan Burton for the synthesis of some starting materials. Supporting Information Available: Detailed procedures and full characterization of new compounds. This material is available free of charge via the Internet at http://pubs.acs.org. OM7004518